Unprecedented Large Hyperpolarizability of Twisted Chromophores in

Jun 18, 2018 - ... degli Studi di Perugia , Via Elce di Sotto, 8 I-06123 Perugia , Italy ... TICT chromophores with interaryl torsional angles in the ...
1 downloads 0 Views 1MB Size
Subscriber access provided by CAL STATE UNIV SAN FRANCISCO

Article

Unprecedented Large Hyperpolarizability of Twisted Chromo-phores in Polar Media Alexander J.-T. Lou, Stefania Righetto, Christopher Jeffrey Barger, Cristiano Zuccaccia, Elena Cariati, Alceo Macchioni, and Tobin J. Marks J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b04320 • Publication Date (Web): 18 Jun 2018 Downloaded from http://pubs.acs.org on June 18, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Unprecedented Large Hyperpolarizability of Twisted Chromophores in Polar Media Alexander J.-T. Lou§, Stefania Righetto†, Christopher Barger§, Cristiano Zuccaccia‡*, Elena Cariati†*, Alceo Macchioni‡*, Tobin J. Marks§* §

Department of Chemistry and the Materials Research Center, Northwestern University, 2145 Sheridan Rd. † Evanston, Illinois 60208, United States. Dipartimento di Chimica dell’Università di Milano and Unità di Ricerca ‡ dell’INSTM di Milano, Via Golgi 19, I-20133 Milano, Italy. Dipartimento di Chimica, Biologia e Biotecnologie and CIRCC, Università degli Studi di Perugia, Via Elce di Sotto, 8 I-06123 Perugia, Italy

Supporting Information Placeholder ABSTRACT: Twisted intramolecular charge transfer (TICT) chromophores exhibit uniquely large second order

optical nonlinearities (µβ). However, their promise as electro-optic (E-O) materials is yet untapped, reflecting a strong tendency to aggregate in low-polarity media, leading to a dramatic fall in µβ. Until now, TICT chromophores in de-aggregating polar solvents suffered decreased response from polarity driven changes in electronic structure. Here we report a new series of benzimidazolium-based TICT chromophores with inter-aryl torsional angles in the range of 64-77°. The most twisted, B2TMC-2, exhibits a large µβvec = -26,000 × 10-48 esu (at 1907 nm) in dilute non-polar CH2Cl2 solution, which is maintained in polar DMF (µβvec= -20,370 × 10-48 esu) as measured by DC electric-field induced second harmonic generation (EFISH). Sterically enforced inter-aryl torsional angles are confirmed by single crystal X-ray diffraction and solution phase Nuclear Overhauser Effect (NOE) NMR, and spectroscopic characterization reveals a zwitterionic/aromatic ground state electronic structure associated with the high µβ. We show that increasingly disrupted conjugation is correlated with increased µβ, even at intermediate twist angles. The excellent performance and reduced aggregation in polar solvents opens new avenues for bridging microscopic and macroscopic chromophore performance.

INTRODUCTION Second order organic nonlinear optical (NLO) materials are of great interest for their ability to generate, process, and switch optical signals for applications such as image reconstruction and optical telecommunication.1,2 In order to create useful optoelectronic devices, the constituent organic chromophores must possess a large molecular hyperpolarizability (β), optical transparency, chemical and thermal stability, and be effectively incorporated into device-ready materials. To date, organic materials have demonstrated excellent electro-optic coefficients in excess of 100 pm/V, as compared to 32 pm/V in the ubiquitous inorganic device material, LiNbO3.1,3-7 Organic molecules have the attraction of being readily tuned structurally and electronically, and chromophores with large β can be rationally designed using structure-property relationships developed over several decades.

In the past, the majority of second order NLO designs have followed a simple paradigm in which the chromophores are composed of donor and acceptor functionalities, bridged by a planar π-system. These “push-pull” systems typically possess a low-lying strongly polarized, charge transfer (CT) state, and can therefore be described by the simple “two-state model”, which provides a qualitative relationship between β and tunable molecular properties (eq. 1).8 Here β is related to the CT energy (Eeg), transition dipole moment (µeg), and the change in state dipole moment (∆µeg = µe - µg).

ߚ=

మ ୼ఓ೐೒ ఓ೐೒ మ ா೐೒

(1)

Another approach, known as bond-length alternation (BLA), manipulates the contributions of CT limiting resonance forms to the ground state electronic structure in order to enhance β.9-11 BLA based chro-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

mophores (see CLD-1, Figure 1) generally contain donor and acceptor groups bridged by extended, planar, polyene backbones. Such structures have been critical to the elucidation of structure-property relationships and can exhibit µβ in excess of 17,600 × 10-48 esu.12 However, these BLA chromophores are often limited by photochemical and oxidative instability and rely on low-lying CT states which erode the optically transparent window in the near IR.12,13 Other approaches such as multi-dimensional chromophores,14,15 modulated conjugation,16 and auxiliary donor/acceptors17 have also afforded promising increases in hyperpolarizability. CN N C5H11

Page 2 of 14

ties to host-guest polymer matrices, where high effective chromophore concentration is required for useful bulk response.5,20 While polar environments can provide stabilizing interactions which disrupt aggregate formation, they also severely depress the NLO response of the aforementioned TICT chromophores. For example, at low concentrations, the maximum µβvec of TMC-3 in polar DMF is about 1/5 of that in DCM (Figure 2).20

RO OR

CN

N

TMC-3

C3H7 C5 H11

C3H7

C4H9

N

N R

NC NC

CN

TMC-2

C5H11

N

CLD-1

R

CN

BXTMC-2 R = H, Me

O NC

CN CN

Figure 1. TICT chromophores TMC-2 and TMC-3, BLA inspired CLD-1, and BXTMC-2 chromophores in this work.

In 1997, Albert et al. challenged the notion that planarity was a prerequisite to large β, suggesting that a twisted intramolecular charge transfer (TICT) chromophore could surpass previous efforts.18 Following this report, TICT chromophore TMC-3 was experimentally shown to exhibit µβvec = -488,000 × 10-48 esu (in dilute DCM solution) (Figure 2),5,18-22 exceeding the previous best µβ/MW by a factor of ~20×, and approaching the fundamental limits on β proposed by Kuzyk et al.1 The exceptional NLO response stems from a twisted bi-aryl bridge fragment and aromatic stabilization of the charge separated ground state, which leads to a low-lying CT (small Eeg) and large ∆µeg.23 Further experimental work on TICT chromophores focused on, (1) the impact of torsional angle;22 (2) adding multiple twisted fragments;21 (3) extending the conjugation.5 Nevertheless, while the NLO response of TICT chromophores is extraordinarily high in dilute, nonpolar solution, these chromophores suffer from dipole driven aggregation, yielding anti-parallel dimers for which µβ = 0.5 This tendency poses a challenge when attempting to translate solution phase proper-

Figure 2. EFISH derived µβvec (solid lines) and PGSE NMR-derived aggregation number (dotted lines) of TMC-3 as a function of concentration in DCM (red diamonds), DMSO-d6 (black dotted line), and DMF (solid black lines). Figure reproduced from Ref. 5.

The desire to translate molecular performance to bulk has been a long-standing goal within the community; past approaches include Langmuir-Blodgett films, polled polymers, layer-by-layer self-assembly, modification of the dielectric environment, and cochromophore incorporation.2,8,24-27 However, as we will discuss herein, structural modification of the chromophores themselves is, in this case, a necessary and productive strategy. In the present study, we substitute the pyridinium acceptor fragment in TMC-2 with a benzimidazolium acceptor, leading to changes in electronic structure and allowing the strategic placement of two sterically demanding alkyl substituents. The steric interaction of these alkyl substituents with orthomethyl groups creates a significant twist angle between the donor and acceptor π-ring planes. The new BXTMC-2 family (Scheme 1) exhibits excellent stability, solubility, and NLO performance. We show that unlike previous TICT generations, these chromophores perform well in polar media, opening routes towards device incorporation which were previously unavailable for TICT chromophores. Furthermore, we characterize the key structural properties of these

2 ACS Paragon Plus Environment

Page 3 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Scheme 1. Synthesis and structures of BXTMC-2 chromophores.

t

*Reagents and conditions: (i) phenylenediamine, 1,4-dioxane, bubbling air at 85° C; (ii) KO Bu, C6H13Br, KI (cat.), t t THF 60° C; (iii) malononitrile, NaO Bu, Pd(PPh3)4, DME, 85° C; (iv) KO Bu then C5H11OTf in acetone at 0° C.

systems and use this information to augment previous observations of NLO response in relation to disrupted conjugation and intrinsic chromophore efficiency. RESULTS Having synthesized a series of benzimidazolium based TICT chromophores with varied expected torsional angles, we investigate their solid state and solution phase properties. Particular attention is paid to changes accompanying increased steric encumbrance, and to the relationship of BXTMC-2 to TMC2. Single crystal X-ray diffraction and NMR spectroscopy are used to determine the molecular geometry, including the bi-aryl torsional angle. Linear optical absorbance in a range of solvents, vibrational spectroscopy, and electrochemical measurements provide a detailed description of the electronic structure and the environmental dependence thereon. Nonlinear optical measurements using DC electric-field induced second harmonic generation (EFISH) are then used to determine µβvec and to assess aggregation tendencies. Experimental details including synthetic and characterization procedures and relevant spectra are included in the Supporting Information. Synthesis of BXTMC-2 chromophores. The synthesis of the BXTMC-2 family (Scheme 1) begins with condensation of phenylene diamine with the desired 4-bromo-benzaldehydes in the presence of air to form intermediates 1A-1C. The acidic proton is then removed, and selective N-alkylation is performed with 1-bromohexane. The aryl bromides in molecules 2A-2C can be efficiently converted to the

corresponding dicyanomethanide functionalities via a Pd-catalyzed coupling with malononitrile. The products of this reaction (3A-3C) appear by NMR to be the zwitterionic isomer, where the acidic malononitrile proton migrates to the benzimidazole ring. As such, it is then necessary to treat structures 3A-3C with base prior to alkylation to yield the final BXTMC-2 products, and to perform alkylation at low temperature to ensure selectivity. The identity and purity of the final products was confirmed by a standard battery of spectroscopic and physical methods as described below and in the SI. BXTMC-2 Solution structures by 1H and 15N NMR. 1H NMR spectra of the BXTMC-2 chromophores in DMSO-d6 (Figure 3) shows two sets of signals in the aromatic region; (1) resonances at δΗ = 7.6 – 8.2 ppm associated with the electron-deficient benzimidazolium acceptor fragment and (2) and signals at δΗ = 6.6 – 7.5 ppm for the electron-rich aryl donor fragment. Hc in DMSO-d6 shifts to higher field with increased bi-aryl torsion, from B0TMC-2 (δΗ = 6.97 ppm) to B1TMC-2 (δΗ ~ 6.82 ppm) and B2TMC-2 (δΗ = 6.68 ppm) indicating an increased electron-richness in that molecular fragment. Both increasing the NOE NMR determined twist angle (see more below) and addition of electron donating methyl groups likely contribute to these shifts. The acceptor peaks (He, Hf) do not appear to be very sensitive to such changes, and exhibit ∆δΗ < 0.1 ppm between B0- and B2TMC-2.

3 ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 14

of relative δ15N values indicates a larger variation of the bi-aryl torsion between B0- and B1TMC-2 than between B1- and B2TMC-2, in agreement with NOE data (see below).

1

Figure 3. H NMR spectra of BXTMC-2 in DMSO-d6: B0TMC-2 (green), B1TMC-2 (blue) and B2TMC-2 (black). Peaks related to Hc are bracketed in red; He and Hf are in purple.

1

15

Figure 4. Left: Superimposed H, N HMBC NMR spectra of BXTMC-2 in DMSO-d6. Right: Trend of δ15N (ppm) as a function of the inverse of the optically determined HOMO-LUMO energy gap (1/ECT).

In contrast to the above 1H aromatic chemical shift trends, the 15N chemical shifts (δ15N) of benzimidazolium resonances (Figure 4, left) are displaced significantly and systematically within the BXTMC-2 series, with δ15N increasing from B0- (-216.8 ppm) to B1- (-214.0 ppm) and B2TMC-2 (-212.7 ppm). The observed δ15N trend is determined by the paramagnetic contribution, as is usual for nuclei heavier than 1 H. Consistently, a good linear correlation is obtained when plotting δ15N versus the inverse of the optically determined HOMO-LUMO energy gap (ECT, Table 2), according to the Ramsey Equation(Figure 4, right).28-30 This δ15N trend also reasonably reflects increasing accumulation of positive charge at the benzimidazolium nitrogen atoms on progressing from B0TMC-2 to B2TMC-2,31-33 and indicates an increasing contribution of the aromatic resonance form to the ground-state structure.34,35 Comparison

Crystallographic Characterization of BXTMC-2 Chromophores. Single crystal diffraction characterization provides insight both into the molecular geometries of the chromophores, as well as the modes of solid state intermolecular interactions operative (Figures 5 and 6). Large bi-aryl torsional angles of 70o, 89o, 79o are observed for B0-, B1-, and B2TMC-2 respectively, confirming the steric contribution of both H/CH2 and CH3/CH2 repulsions. Two key bond distances are also examined (Figure 5, Table 1). First, the bridging (ring)C-C(ring) bond distance is 1.465(4), 1.474(8), and 1.478(4) Å for B0-, B1-, and B2TMC-2, respectively. These bond lengths are more similar to those observed for bi-mesitylene (1.501 Å)36 and bi-aryls (1.487 Å)37 than typical C=C bonds (~1.33 Å),38 indicating that these molecules are best described by linked aromatic structures rather than as quinones. The comparison of B0- and B2TMC-2 structures indicates that the increased torsion arising from added methyl groups leads to a slight elongation of the (ring)C-C(ring) bond, likely due to a decrease in its double bond character. The uncertainty in B1TMC-2 bond distances precludes comparison with the other structures. The second key bond distance, (ring)C-C(CN)2, is found to be 1.445(4), 1.438(8), and 1.450(4) Å for B0-, B1-, and B2TMC-2 respectively (Table 1). These values are longer the analogous bond lengths found in quinoidal TCNQ (1.373 Å),39 suggesting that the dicyanomethanide group supports significant negative charge. The increase in bond length between B1- and B2TMC-2 indicates a decrease in double bond nature of the ring(C)-C(CN)2 bond, consistent with increased aromatic character of the structure (see Figure 11). The similarity of B0- and B1TMC-2 ring(C)C(CN)2 distances may be a coincidental result of different packing structures; BoTMC-2 exhibits close contact between donor and acceptor of adjacent molecules potentially stabilizing the negative charge.

Figure 5. ORTEP drawing of chromophore B2TMC-2 with 30% probability ellipsoids showing key bond metrical parameters.

4 ACS Paragon Plus Environment

Page 5 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

a driving force in these interactions. There is no evidence of co-crystalized solvent associated with these structures, and the unit cells lack solvent accessible voids. Scheme 2. BXTMC-2 chromophore structures and labeling scheme showing syn and anti configurations. R

He Ha

He H a

Ha

N Me

Hc

N Me Ha He Ha R

CN CN He

B2TMC-2 anti He Hb

R

N He

Ha R

Hd Hb

B1TMC-2 anti

N

Hd

N

Hd Ha

Ha R

He H b Hc

Ha Hc

CN CN

B0TMC-2 anti

Ha

N Me

R

R

Ha

N Me

CN CN

N He

Hb R

Hc

Hd Ha

CN CN

B1TMC-2 syn

Figure 6. Crystal structures of the BXTMC-2 chromophore series. Left: packing of the tetrameric centrosymmetric unit cells (alkyl groups excluded for clarity). Right: Closest packed dimeric structures with relevant packing distances measured in Å. Table 1. Crystallographic bond and intermolecular packing distances

Distance (Å) Packing

(ring)CC(ring)

(ring)CC(CN)2

distance

B0TMC-2

1.465(4)

1.445(4)

9.225(6)

B1TMC-2

1.474(8)

1.438(8)

7.636(8)

B2TMC-2

1.478(4)

1.450(4)

7.175(4)

The BXTMC-2 crystal structures all exhibit antiparallel dimer packing within the centrosymmetric tetramers which constitutes the unit cell (Figure 6). The distance between the bridging carbon on the donor ring of the anti-parallel aligned pairs provides an estimate of the closeness of the intermolecular interaction. This distance contracts on progressing from BoTMC-2 (9.220(6) Å) to B1TMC-2 (7.636(8) Å) and B2TMC-2 (7.175(4) Å), suggesting stronger intermolecular interactions, particularly in B2TMC2. This trend is in agreement with a computed increase in ground state dipole moment, which is likely

Figure 7. Computed average distance for B0TMC2 (Hd-Ha), B1TMC-2 (Me-Ha and Hd-Hb) and B2TMC-2 (Me-Ha) chromophores in the solid state as a function of dihedral twist angle (θ). The cross points indicate the experimentally NOE-derived distances and the corresponding average twist angle in solution.

Nuclear Overhauser Effect (NOE) NMR Characterization of Chromophore Structures in Solution. 1H-1H NOE NMR experiments, which measure through space dipolar coupling, were undertaken in DMSO-d6 to estimate the average twist angle in solution for the BXTMC-2 chromophore series. For B2TMC-2, NOE measurements were carried out by irradiating the Me singlet (δH = 1.93 ppm at 298K) and measuring NOEs at both Ha and Hc (Scheme 2). For B0TMC-2, the Hd doublet (δH = 7.44 ppm at 298K) was irradiated and NOEs were measured on

5 ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Ha. In the case of B1TMC-2 both the Me singlet (δH = 2.02 ppm at 298K) and the Hd doublet (δH = 7.31 ppm at 298K) were irradiated, and the following NOEs were quantified: Me-Hc, Me-Ha, and Hd-Hb. From the measured NOEs, average values of cross relaxation rate constants ()40 at different temperatures (Figure S1) and average internuclear distances 41 were obtained (Table S4) using the methodology described previously.5 In order to correlate the average internuclear distance measured in solution between aromatic (Me or Hd groups) and benzimidazole (Ha and Hb protons) moieties with the average twist angle (θ) in solution, average internuclear distances were computed, under rIS-6 averaging,42 for seven static conformations (θ approximately equal to 60°, 65°, 70°, 75°, 80°, 85° and 90°) starting from the solid state X-ray structures of the BXTMC-2 series. Considering the simplified approach detailed in the Supporting Information, the computed average distances, reported in Figure 5, are somewhat biased toward shorter distances (i.e., toward the distance of closest approach); consequently, the derived twist angles in solution (cross points in Figure 5) are somewhat underestimated, and should be better considered as a lower limit for θ. There is no doubt that B0TMC-2 exhibits the smallest twist angle (64°) in solution (Figure 7). However, B1- and B2TMC-2 are found to have very similar twist angles (77-78° and 76°, respectively), according to the present analysis. From this analysis, it is clear that both Me/CH2 and H/CH2 interactions play a significant role in bi-aryl torsion. The similarity of the B1- and B2TMC-2 torsional angles is addressed in the discussion section. Infrared (FTIR) Vibrational Spectroscopy. BXTMC-2 compounds were characterized by FTIR both as solids and in solution. The key feature of these spectra is the characteristic C≡N stretching, v(C≡N), with a lower energy side component, presumably the symmetrically and anti-symmetrically coupled modes (Figure 8). In the solid state, B1- and B2TMC-2 exhibit v(C≡N) = (2168, 2133 cm-1) while B0TMC-2 exhibits a slightly lower energy side band at 2128 cm-1. The energy and splitting of the bands is similar to a phenyl malononitrile anion (v(C≡N) = 2163, 2117 cm-1),43 indicating that the dicyanomethanide groups in BXTMC-2 bear a large amount of electron density. In DCM solution, all BXTMC-2 chromophore molecules display significantly higher energy v(C≡N) stretching modes than as solids, and v(C≡N) falls in energy from B0- (2176 cm-1) to B1- (2174 cm-1) and B2TMC-2 (2172 cm-1). This trend indicates a slight

Page 6 of 14

increase in electron density on the dicyanomethanide group accompanying the addition of methyl groups. Similar results were obtained in DMF, with the v(C≡N) band increasing from 2170 cm-1 in B2TMC-2 to 2172 cm-1 in B0TMC-2 (Figure S10). The side band is higher in energy in solution than in the solid state (2139 cm-1 in DCM, 2135-2137 cm-1 in DMF). These observations suggest that the BXTMC2 chromophores exhibit predominantly zwitterionic character which increases in the order DCM < DMF < solid.

Figure 8. FTIR vibrational spectra of B0TMC-2 (green), B1TMC-2 (blue), and B2TMC-2 (black); (A) solid (ATR); (B) DCM solution.

Linear Optical Absorption Spectroscopy as a Function of Solvent. The solution optical absorption of the BXTMC-2 family (Figure 9A) all exhibit a broad low-lying charge transfer (CT) absorption (434 nm – 408 nm) and a higher energy transition (315 nm – 328 nm). The intensity of the CT peak decreases from ε = 24,928 M-1cm-1 in BoTMC-2 to ε = 8156 M1 cm-1 in B2TMC-2. Such reduction in transition intensity suggests increasingly disrupted conjugation, which is in agreement with reduced NOE and crystallographic torsional angles for B0TMC-2, but not for B1 and B2TMC-2. This discrepancy is addressed in detail in the Discussion Section. By integrating the CT peaks (see Supporting Information), one can also extract transition moments (µeg), which reveal

6 ACS Paragon Plus Environment

Page 7 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

relatively large transition strengths of 5.45 D in B0TMC-2 and 3.43 D in B2TMC-2 as compared to Table 2. BXTMC-2 optical absorption and electrochemical data.a f

λCT (nm)

ECT (DCM )

B0TMC-2

E (ACN)

g

(eV)

∆λ h (nm)

Eox (V)

Ered (V)

434

2.55

-64

0.24

-1.94

5.45

21.2

-12.3

B1TMC-2

415

2.62

-65

0.36

-2.03

4.72

23.5

-13.0

B2TMC-2

408

2.67

-57

0.34

-2.10

3.43

24.6

-13.7

556

2.18

~-180

0.39

-1.56

2.10

29.8

-21.4

e

TMC-2 a

µeg(D)

d

µg(D)

b

∆µeg(D)

c

b

All values reported are measured in DCM solution except where noted otherwise. calculated using CAMc d B3LYP/6-31G**. extracted from linear absorption using the McRae equation, see SI for details. extracted from line f ear absorption, see SI for details. data from ref. 5. optical HOMO-LUMO gap estimated from onset of CT absorpg h tion measured from signal onset Solvent shift from CHCl3 to MeOH

2.1 D in TMC-2 (Table 2). The fall in transition intensity is accompanied by a ~27 nm hypsochromic shift from B0TMC-2 to B1TMC-2 to B2TMC-2 in DCM. The higher energy excitation, attributed to a subfragment transition, also undergoes a ~13 nm bathochromic shift, along with a slight increase in transition strength from B0TMC-2 to B2TMC-2. The behavior in Figure 9A is typical of TICT chromophores; as conjugation is disrupted, both sub-fragments of the molecule become isolated and their own characteristic transitions begin to dominate.22

UV-Vis absorption spectra of neat films (Figure S3) are found in all cases to be similar to those in acetone solution, demonstrating that the molecules behave comparably as solids and in relatively polar solutions.

Solvent-dependent spectra (Figure 9B) reveal negative solvatochromic shifts of the CT band, meaning that CT wavelength (λCT) increases with solvent polarity (the shift between MeOH and CHCl3 is reported as ∆λ in Table 2). Such shifts indicate that the ground state dipole moment is larger than that of the first excited state (∆µeg < 0), and therefore is stabilized by the increase in solvent polarity. The value of λCT is determined by Gaussian fitting of the CT peak, which eliminates the effect of the sub-fragment transition on the CT peak position (Table 2). The McRae equation then provides a simple means to calculate ∆µeg by plotting the change in ECT in various solvents (Figure S4) against a solvent polarity function f(ε, n).44-46 v – v0 = -m × f(ε, n2)

(2)

Here n is refractive index and ε is the solvent dielectric constant, and v - v0 is the difference between the CT frequency (in cm-1) in a given solvent and in vacuum. By relating the slope, m, to ∆µeg, an increase in ∆µeg from -12.3 D in B0TMC-2 to -13.7 D in B2TMC-2 is calculated (Table 2). More details of this analysis can be found in the Supporting Information.

Figure 9. UV-Vis linear optical absorption spectra of BXTMC-2 chromophores. (A) BXTMC-2 chromophores in DCM solution; (B) B0TMC-2 in the indicated solvents and as film.

BXTMC-2 Cyclic Voltammetry. Chromophores B1- and B2TMC-2 exhibit reversible reduction at 2.03 and -2.10 V, respectively, while the reduction of B0TMC-2 is irreversible at -1.94 V (Table 2). Oxidation is irreversible for all chromophores, with an onset between 0.24 – 0.36 V. From the oxidation and

7 ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

reduction potentials, the HOMO-LUMO gaps are estimated to be 2.18, 2.39, and 2.44 eV for B0-, B1-, and B2TMC-2, respectively; on average ~0.3 eV lower than the optical ECT, possibly reflecting some mixed character in the optical transitions and solvation reorganization in redox processes.47 NLO Measurements. Measurements of μβvec, the product of the chromophore dipole moment (μ) and the vector part of the molecular first-order hyperpolarizability βvec tensor along the μ direction, were performed by the solution-phase DC EFISH method, which provides direct information on the intrinsic molecular nonlinear optical (NLO) response via eq 3. γEFISH = (μβ/5kT) + γ(−2ω; ω, ω, 0)

(3)

Here, μβ/5kT is the dipolar orientational contribution, and γ(−2ω;ω,ω,0), the third-order term at frequency ω of the incident light, is the electronic contribution to γEFISH, which is negligible for molecules of the type investigated here.48 EFISH measurements at 1907 nm were performed over a concentration range in DCM and DMF to determine, (1) trends in the monomeric response of the BXTMC-2 chromophores in highly dilute solutions; (2) the dependence on solvent environment; (3) the effect of aggregation in different solvent environments. It is well known that highly polar TICT chromophores have a strong tendency to aggregate in solution,5 so the most dilute reliable measurement for each chromophore was taken to be representative of monomeric µβvec. Note that this value reflects the lower bound of µβvec, particularly in the case of B1TMC-2 for which larger values were recorded at higher dilution but with greater uncertainty. Measurements in DCM reveal very large monomeric µβvec= -26,000×10-48 esu for B2TMC-2, and µβvec of -11,730 and -10,300×10-48 esu for B1TMC2 and B0TMC-2 respectively (Figure 10A). The response of B2TMC-2 is on the order of the monomeric response of TMC-2 (Figure 1; µβvec=-24,000×10-48 esu), despite increases in ECT and a reduced ∆µeg. EFISH measurements on B2TMC-2 solutions in polar DMF (Figure 10B) reveal an NLO response of µβvec= -20,370×10-48 esu, which is similar to the measured value in DCM, and about 4× higher than TMC-2 (-5,620×10-48 esu) under the same conditions. B1TMC-2 shows a similarly high value of µβvec = 12,740×10-48 esu in DMF compared to -11,730×10-48 esu in DCM (Table 3). These observations are in sharp contrast to all previous TICT chromophores, which suffer massively decreased NLO responses in polar solvents (Figures 1 and 10B).

Page 8 of 14

In DCM solutions, a significant decrease in µβvec is observed with increasing concentration in the range of 10-5 – 10-4 M for B1- and B2TMC2, and 10-4 – 10-3 M for B0TMC-2. This behavior is consistent with the formation of centrosymmetric aggregates (for which µβ = 0), and mirrors previous observations on TMC2 in DCM (Figure 10). The concentration dependence of NLO response is more pronounced in the most twisted chromophore (B2TMC-2) than in the least twisted (B0TMC-2), likely reflecting the dipole moment enhancement which accompanies bi-aryl torsion. The use of more polar DMF appears to mitigate these effects, and shifts the decreases in NLO response towards higher concentrations (Figure 10B). Table 3. EFISH results for BXTMC-2 chromophores in DCM and DMF at 1907 nm

µβ (esu × 10-48)a

β (esu × 10-30)b

DCM

DMF

DCM

DMF

B0TMC-2

-10,300

-8,400

-486

-396

B1TMC-2

-11,730

-12,740

-499

-542

B2TMC-2

-26,000

-20,370

-1,056

-828

-24,000

-5,620

-805

-189

c

TMC-2 a

Highest dilution with reliable data is used to estimate b calculated using DFT derived dipole moments c in Table 1. data from ref. 5.

µβvec.

Figure 10. EFISH measurements of BXTMC-2 chromophores at 1907 nm (A) in DCM, with TMC-2 data from

8 ACS Paragon Plus Environment

Page 9 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

ref. 5; (B) comparison of µβ in DCM (solid lines) and DMF (dashed lines) for B2TMC-2 and TMC-2.

DISCUSSION. Synthesis of the BXTMC-2 Family. The synthetic route in Scheme 1 used to create BXTMC-2 chromophores offers some significant advantages over the previous generations; (1) the final product here can be obtained in only 4 steps; (2) there is an opportunity to introduce asymmetric or functionalized sidechains; (3) introduction of the steric bulk in two separate steps leads to high yields and a further degree of synthetic flexibility. By forming the benzimidazolium acceptor group via annulation rather than coupling, we avoid repeated use of aryl halide functionality, significantly reducing the number of required synthetic steps. Furthermore, BXTMC-2 family exhibits excellent thermal stability, with thermogravimetric analysis derived decomposition temperatures in excess of 310 °C (Figure S5). BXTMC-2 Electronic Structure. The ground state electronic structure of BXTMC-2 chromophores can be represented as a mixture of zwitterionic/aromatic (ZA) and quinoidal/neutral (NQ) resonance forms (Figure 11). As shown in the Results section, all three R

C 5H 11 R

NC

N

NC

NC

N

NC

R'

C 4H 9

Aromatic/ zwitterionic (ZA)

C 5 H11

While both crystallographic and NOE characterization data definitively show that Me/CH2 interactions lead to larger twist angles (θ) than do H/CH2, they also indicate similar θ for B1- and B2TMC-2. This is in contrast to the decrease in the measured CT transition moment (µeg) for B2- versus B1TMC-2, which suggests a reduction of conjugation. The reason for this observed discrepancy may relate to the dynamic nature of θ. Although the two chromophores have similar average θ by NOE, they are able to access a range of θ through thermal excitation of the vibrational ring-twisting mode. As added Me/CH2 interactions are shown to increase twisting, it is reasonable that B2TMC-2, with twice the steric resistance, is less able to access smaller θ values than B1TMC-2 at a given temperature. So, despite having similar average θ, thermal energy more easily populates less twisted configurations with large µeg for B1TMC-2, leading to larger observed CT absorption than for B2TMC-2. It should be noted that quantum computation indicates that the BXTMC-2 chromophores will exhibit similar θ values in DCM and DMSO (Tables S9 and S10), and so we have successfully accessed intermediate θ in polar and nonpolar solution which are less than 80°, but still deviate significantly from planarity.

N N R'

C 4 H9

Quinoidal/ neutral (NQ)

Figure 11. Quinoidal and aromatic contributions to the BXTMC-2 electronic structure.

BXTMC-2 chromophores are best described by the ZA structure. The crystallographic (ring)C-C(ring) bond lengths are similar to those found in bi-aryls and bi-mesitylene, and vibrational frequencies of the C(CN)2 group show that the donor group bears an essentially full negative charge. This is consistent with previous work by Isborn et al., which notes that even at modest twist angles, aromatic stabilization of the donor and acceptor imparts dominant ZA character in similar molecules.49 However, there is also evidence here supporting a subtle increase in ZA character from B0- to B1- and B2TMC-2. Shifts of the 1H and 15N NMR resonance positions, increased solvatochromic shifts, lengthening of crystallographically derived (ring)C-C(ring) bond distances, and increased v(C≡N) energies support the trend in ZA character increasing in the progression, B0- < B1- < B2TMC-2.

Figure 12. HOMO and LUMO contours of BXTMC-2 chromophores with 40° and 90° twist angles.

Comparison to TMC-2 and Previous TICT Chromophores. Comparison to other TICT chromophores sheds light on the impact of the benzimidazolium group as compared to other acceptors previously employed. The transition strengths, even in B2TMC-2 (ε = 8156 M-1cm-1) are markedly greater than those reported for TMC-2 (ε = 1840 M-1cm-1) and TMC-3 (ε = 2090 M-1cm-1), likely reflecting the reduced torsional angles. The enhanced transition moment may also relate to the change in acceptor group, which dictates the attributes of the LUMO. Electronic structure calculations show that a large

9 ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

portion of the LUMO electron density is located on the bridging carbon, yielding a strong overlap with HOMO located on the donor fragment, particularly at intermediate torsional angles (Figure 12). These calculations also suggest that increased torsion largely manifests as localization of the LUMO on the benzimidazolium fragment, leading to larger ∆µeg (Tables S7 and S8). The BXTMC-2 CT transition in the 407-434 nm range is at much higher energy than that of previous twisted chromophores (λCT = 569 nm for TMC-2), due to the ~0.5 eV increase of the LUMO level as compared to TMC-2. This constitutes a significant advantage; a device utilizing BXTMC-2 chromophores could operate without linear loss in a far broader spectral window. In fact, the absorption edge in polar solvents is close to the visible region cutoff, making a high performance photonic device material with visible region transparency a distinct possibility. Despite relatively large ground state dipole moments, the BXTMC-2 chromophores are less sensitive to solvent environment than any previous second order NLO TICT chromophore. The maximum measured solvatochromic shift (measured from CHCl3 to MeOH) of B2TMC-2 is -57 nm, as compared to ~-180 nm for TMC-2, showing that replacement of the pyridinium fragment with benzimidazolium significantly reduces the dependence of the relative energy level spacing on the dielectric environment. Several factors may contribute to this observation: (1) the positioning of the bulky alkyl groups may reduce solvent interactions with the acceptor moiety; (2) the reduced CT distance versus TMC-2 decreases the dipole moment and therefore the importance of solvent stabilization; (3) the benzimidazolium acceptor provides more effective stabilization of the ground state positive charge than previous acceptor groups. Monomeric NLO Response in DCM and DMF. The contribution of both CH2/H and CH2/CH3 steric interactions to the bi-aryl torsion allows access to intermediate solution twist angles which were not available in previous studies. The present evidence suggests that twist angle is directly related to µβvec, and that the highest nonlinear response is achieved above 70°. The justification for large NLO response of this nature is now well-known to arise from the ZA/NQ balance in the molecule.19,50 Twisting, or otherwise disrupting conjugation, serves to simultaneously destabilize the ground state and create very large charge separation leading to giant dipole moments. The large NLO response is a consequence of these factors, as shown by extensive experimental

Page 10 of 14

and computational work. This concept has also been demonstrated by recent theoretical work showing enhanced β with disrupted conjugation, as provided by large torsional angles.16,51 In the past, TICT chromophores performed poorly in polar environments, despite marked decreases in aggregation.5,22 For TMC-2, NLO response was shown both experimentally and computationally to peak in moderately polar solvents, then diminish with increasing polarity. This behavior was attributed to the stabilization of the zwitterionic ground state relative to the excited state manifold in highly polar environments.19,52,53 In the case of B2TMC-2, the similarity of NLO response observed in DMF and DCM is likely a result of reduced dependence of Eeg on solvent polarity, as evidenced by relatively weak solvatochromic shifts (∆λ in Table 2). Of the TICT chromophores characterized to date, B2TMC-2 shows the highest reported nonlinearity in DMF at 10-3 M (the highest measured concentration), highlighting the newfound ability to benefit from polar environments. It should be noted that according to the twolevel model, changes in the linear absorption should lead to decreased NLO response in DMF versus DCM. However, it was previously noted that the twolevel model does not always properly capture the impact of the dielectric medium.27 Concentration-Dependence of NLO Response. In DCM, the aggregation tendencies of B2TMC-2 and TMC-2 are similar; high responses observed at low concentrations quickly diminish in the range of 10-4 – 10-5 M. The observed aggregation tendencies are consistent with the tightly packed centrosymmetric unit cells in the crystal structures of B1TMC-2 and B2TMC-2. The relatively short linear alkyl groups used here do not prevent interactions between the donor group and the face of the benzimidazolium fragment, as evidenced by the crystallographic packing. However, the design of BXTMC-2 allows for facile modification of the sidechains; adding branched or dendritic groups has proven effective in the past, and may be used to prevent close access to the face of the acceptor moiety.22,54 Aggregation of B2TMC-2 is less favorable in DMF, as shown by the extension of high NLO response to concentrations in the range of 10-3 – 10-4 M. Such improvements are encouraging, but clearly must be supplemented by the aforementioned chemical modification of the sidechains. Should this strategy prove effective in mitigating aggregation, it could be combined with the use of polar matrices, which had not been compatible with large NLO response in TICT chromophores until this point.

10 ACS Paragon Plus Environment

Page 11 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Intrinsic NLO Response. In order to compare systems of different size and properties, it is useful to examine the intrinsic hyperpolarizability defined by Kuzyk as βint = (observed β)/(theoretical maximum, βmax), for a structure with a particular number of polarizable electrons and a characteristic low-lying optical transition.1 Here, βmax is derived from linear absorbance measurements as described in Section S2 of the Supporting Information. B2TMC-2 yields a remarkably high βint = 0.37, which is one of the largest values ever reported in the open literature (Table 4). Even modestly twisted B0TMC-2 exhibits a large βint = 0.14, still remarkably higher than the vast majority of efforts which lie below 10-3/2. Thus, the BXTMC-2 chromophore series has proven capable of large NLO response arising from the efficient application of a small number of electrons. Table 4. Computed theoretical (β βmax) and intrinsic (β β int), and EFISH (β β EFISH) hyperpolarizability in DCM at 1907 nm

βEFISH

β (× 10-30 esu) βmax

βint

B0TMC-2 B1TMC-2

-486 -499

-3370 -3070

0.14 0.16

B2TMC-2

-1056

-2870

0.37

Conclusions. A new series of twisted π-system electro-optic chromophores was synthesized by expeditious new routes and shown to be qualitatively similar to previous TICT generations, with large ground state dipole moments, low-lying CT transitions, and primarily zwitterionic/aromatic character. The same theories which justified NLO response in the past also apply here, as evidenced by the dependence of µβ on twist angle. The introduction of steric hindrance using a benzimidazole moiety permits access to intermediate torsional angles and demonstrates the effect of experimentally twisting in this range for the first time. The highlight of the present work is the strong NLO performance of chromophore B2TMC-2 in DMF, which represents a 5× increase over archetypical chromophore TMC-2. We recognize that it is now possible to perform EFISH and E-O measurements in highly polar environment, opening new pathways to address detrimental aggregation. We are now pursuing promising directions involving modification of solvent polarity through addition of organic salts, means of artificially increasing polymer matrix polarity, and introduction of dendritic sidechains. The use of highly polar matrices may be a key factor in

preventing aggregate formation and realizing the exciting full potential of TICT chromophores. ASSOCIATED CONTENT Supporting Information is available free of charge on the ACS Publications website. Detailed characterization of BXTMC-2 is available as well as crystallographic information and experimental procedures for the synthesis of all structures.

AUTHOR INFORMATION Corresponding Author

*[email protected] Funding Sources

No competing financial interests have been declared.

ACKNOWLEDGMENT This work was supported by AFOSR MURI grant FA9550_14_1_0040, and made use of the IMSERC facility at Northwestern University, which has received support from the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF ECCS-1542205); the State of Illinois and International Institute for Nanotechnology (IIN). A. J.-T. L. thanks NDSEG for a Graduate Research Fellowship. A. M. and C. Z. thank PRIN 2015 (20154X9ATP 004) for financial support. The University of Perugia and MIUR are also acknowledged for financial support to the project AMIS, through the program “Dipartimenti di Eccellenza – 2018-2022”. We also thank Dr. C. Malliakas for assistance in solving the crystal structure of BoTMC-2.

REFERENCES (1) Kuzyk, M. G.; Singer, K. D.; Stegeman, G. I. Theory of Molecular Nonlinear Optics. Adv. Opt. Photonics 2013, 5, 4. (2) Benight, S. J.; Bale, D. H.; Olbricht, B. C.; Dalton, L. R. Organic Electro-Optics: Understanding Material Structure/Function Relationships and Device Fabrication Issues. J. Mater. Chem. 2009, 19, 7466-7475. (3) S.-S. Sun; Dalton, L. R.: Introduction to Organic Electronic and Optoelectronic Materials and Devices; Taylor & Francis: New York, 2008. (4) Cho, M. J.; Choi, D. H.; Sullivan, P. A.; Akelaitis, A. J. P.; Dalton, L. R. Recent Progress in Second-Order Nonlinear Optical Polymers and Dendrimers. Prog. in Polym. Sci. 2008, 33, 10131058. (5) Kang, H.; Facchetti, A.; Jiang, H.; Cariati, E.; Righetto, S.; Ugo, R.; Zuccaccia, C.; Macchioni, A.; Stern, C. L.; Marks, T. J. Ultra-large Hyperpolarizability Twisted π-electron System Electro-Optic Chromophores. J. Am. Chem. Soc. 2007, 129, 32673286. (6) Kim, T.-D.; Luo, J.; Cheng, Y.-J.; Shi, Z.; Hau, S.; Jang, S.-H.; Zhou, X.-H.; Tian, Y.; Polishak, B.; Huang, S.; Ma, H.; Dalton, L.; Jen, A. K.-Y. Binary Chromophore Systems in Nonlinear Optical Dendrimers and Polymers for Large Electrooptic Activities. J. Phys. Chem. C 2008, 112, 8091-9098.

11 ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(7) Cheng, Y.-J.; Luo, J.; Huang, S.; Zhou, X.; Shi, Z.; Kim, T.D.; Bale, D. H.; Takahashi, S.; Yick, A.; Steier, W.; Jen, A. K.-Y. Donor-Acceptor Thiolated Polyenic Chromophores Exhibiting Large Optical Nonlinearity and Excellent Photostability. Chem. Mater. 2008, 20, 5047-5054. (8) Prasad, P. N.; Williams, D. J.: Introduction to Nonlinear Optical Effects in Molecules and Polymers; John Wiley & Sons: New York, 1991. (9) Meyers, F.; Marder, S. R.; Pierce, B. M.; Bredas, J. L. Electric Field Modulated Nonlinear Optical Properties of Donor Acceptor Polyenes: Sum-Over-States Investigation of the Relationship Between Molecular Polarizabiltiesi (α, β, and γ) and Bond Length Alternation. J. Am. Chem. Soc. 1994, 116, 1070310714. (10) Marder, S. R.; Kippelen, B.; Jen, A. K. Y.; Peyghambarian, N. Design and Synthesis of Chromophores and Polymers for Electro-Optic and Photorefractive Applications. Nature 1997, 388, 845. (11) Gorman, C.; Marder, S. R. Effect of Molecular Polarization on Bond-Length Alternation, Linear Polarizatbiilty, First and Second Hyperpolarizability in Donor-Acceptor Polyenes as a Function of Chain Length. Chem. Mater. 1995, 7, 215-220. (12) Zhang, C.; Dalton, L.; Oh, M.-C.; Zhang, H.; Steier, W. Low Vπ Electrooptic Modulators from CLD-1: Chromophore Design and Synthesis, Material Processing, and Characterization. Chem. Mater. 2001, 13, 3043-3050. (13) Galvan-Gonzalez, A.; Belfield, K. D.; Stegeman, G. I.; Canva, M.; Marder, S. R.; Staub, K.; Levina, G.; Twieg, R. J. Photodegradation of Selected π-conjugated Electro-Optic Chromophores. J. Appl. Phys. 2003, 94, 756-763. (14) Argouarch, G.; Veillard, R.; Roisnel, T.; Amar, A.; Boucekkine, A.; Singh, A.; Ledoux, I.; Paul, F. Donor-Substituted Triaryl-1,3,5-triazinanes-2,4,6-triones: Octupolar NLO-phores with a Remarkable Transparency–Nonlinearity Trade-off. New J. Chem. 2011, 35, 2409. (15) Kang, H.; Zhu, P.; Yang, Y.; Facchetti, A.; Marks, T. J. SelfAssembled Electrooptic Thin Films with Remarkably BlueShifted Optical Absorption Based on an X-Shaped Chromophore. J. Am. Chem. Soc. 2004, 126, 15974-15975. (16) Perez-Moreno, J.; Zhao, Y.; Clays, K.; Kuzyk, M. G. Modulated Conjugation as a Means for Attaining a Record High Intrinsic Hyperpolarizability. Opt. Lett. 2007, 32, 59-61. (17) Leclercq, A.; Zojer, E.; Jang, S. H.; Barlow, S.; Geskin, V.; Jen, A. K.; Marder, S. R.; Bredas, J. L. Quantum-Chemical Investigation of Second-order Nonlinear Optical Chromophores: Comparison of Strong Nitrile-Based Acceptor End Groups and Role of Auxiliary Donors and Acceptors. J. Chem. Phys. 2006, 124, 044510. (18) Albert, I. D. L.; Marks, T. J.; Ratner, M. A. Conformationally-Induced Geometric Electron Localization. Interrupted Conjugation, Very Large Hyperpolarizabilities, and Sizale Infrared Absorption in Simple Twisted Molecular Chromophores. J. Am. Chem. Soc. 1997, 119, 3155-3156. (19) Brown, E. C.; Ratner, M. A.; Marks, T. J. Nonlinear Response Properties of Ultralarge Hyperpolarizability Twisted Pi-System Donor-Accpetor Chromophores. J. Phys. Chem. 2007, 112, 44-50. (20) Kang, H.; Facchetti, A.; Stern, C. L.; Rheingold, W. S.; Marks, T. J. Efficient Synthesis and Structural Characteristics of Zwitterionic Twisted π-electron System Biaryls. Org. Lett. 2005, 7, 3721. (21) Shi, Y.; Frattarelli, D.; Watanabe, N.; Facchetti, A.; Cariati, E.; Righetto, S.; Tordin, E.; Zuccaccia, C.; Macchioni, A.; Wegener, S. L.; Stern, C. L.; Ratner, M. A.; Marks, T. J. UltraHigh-Response, Multiply Twisted Electro-optic Chromophores: Influence of π-System Elongation and Interplanar Torsion on Hyperpolarizability. J. Am. Chem. Soc. 2015, 137, 12521-12538.

Page 12 of 14

(22) Wang, Y.; Frattarelli, D. L.; Facchetti, A.; Cariati, E.; Tordin, E.; Ugo, R.; Zuccaccia, C.; Macchioni, A.; Wegener, S. L.; Stern, C. L.; Ratner, M. A.; Marks, T. J. Twisted π-Electron System Electrooptic Chromophores. Structural and Electronic Consequences of Relaxing Twist-Inducing Nonbonded Repulsions. J. Phys. Chem. C 2008, 112, 8005-8015. (23) Albert, I. D. L.; Marks, T. J.; Ratner, M. A. Remarkable NLO Response and Infrared Absorption in Simple Twisted Molecular π-Chromophores. J. Am. Chem. Soc 1998, 120, 1117411181. (24) Kanis, D. R.; Ratner, M. A.; Marks, T. J. Design and Construction of Molecular Assemblies with Large 2nd-Order Optical Nonlinearities - Quantum-Chemical Aspects. Chem. Rev. 1994, 94, 195-242. (25) Tabor, C. E.; Dalton, L. R.; Kajzar, F.; Kaino, T.; Koike, Y. Theory-Guided Nano-Engineering of Organic Electro-Optic Materials for Hybrid Silicon Photonic, Plasmonic, and Metamaterial Devices. Proc. SPIE 2013, 8622, 86220J. (26) Liao, Y.; Anderson, C. A.; Sullivan, P.; Akelaitis, A. J. P.; Robinson, B. H.; Dalton, L. Electro-Optical Properties of Polymers Containing Alternating Nonlinear Optical Chromophores and Bulky Spacers. Chem. Mater. 2006, 18, 10621067. (27) Bale, D. H.; Eichinger, B. E.; Liang, W.; Li, X.; Dalton, L. R.; Robinson, B. H.; Reid, P. J. Dielectric Dependence of the First Molecular Hyperpolarizability for Electro-optic Chromophores. J. Phys. Chem. B 2011, 115, 3505-3513. (28) Facelli, J. C.: Shielding Calculations: Perturbation Methods. In Encyclopedia of Nuclear Magnetic Resonance; John Wiley: London, 1996. (29) Ramsey, N. F. The Internal Diamagnetic Field Correction in Measurements of the Proton Magnetic Moment. Phys. Rev. 1950, 77, 567. (30) Ramsey, N. F. Magnetic Shielding of Nuclei in Molecules. Phys. Rev. 1950, 78, 699. (31) Chen, B. C.; von Philipsborn, W.; Nagarajan, K. 15N‐NMR Spectra of Azoles with Two Heteroatoms. Helv. Chim. Acta 1983, 66, 1537-1555. (32) Marelius, D. C.; Moore, C. E.; L., A.; Rheingold, A. L.; Grotjahn, D. B. Reactivity Studies of Pincer Bis-protic Nheterocyclic Carbene Complexes of Platinum and Palladium Under Basic Conditions. Beilstein J. Org. Chem. 2016, 12, 13341339. (33) Nieto, C. I.; Cabildo, P.; García, M. A.; Claramunt, R. M.; Alkorta, I.; Elguero, J. An Experimental and Theoretical NMR Study of NH-Benzimidazoles in Solution and in the Solid State: Proton Transfer and Tautomerism. Beilstein J. Org. Chem. 2014, 10, 1620-1629. (34) Witanowski, M.; Stefaniak, L.; Webb, G. A.: Nitrogen NMR Spectroscopy. In Annual Reports on NMR Spectroscopy; Acedemic Press: London, 1993; Vol. 25. (35) Mason, J.: Nitrogen NMR. In Encyclopedia of Nuclear Magnetic Resonance; John Wiley: London, 1996. pp. 3222. (36) Frohlich, R.; Musso, H. Notiz uber die Kristallstruktur des Bimesityls. Chem. Ber. 1985, 118, 4649-4651. (37) Allen, F. H.; Watson, D. G.; Brammer, L.; Orpen, A. G.; Taylor, R.: Typical Interatomic Distances: Organic Compounds. In International Tables for Crystallography, 2006; pp 790-811. (38) Linde, D. R. A Survey of Carbon-Carbon Bond Lengths. Tetrahedron 1962, 17, 125-134. (39) Flandrois, P.; Chasseau, D. Longueurs de Liaison et Transfert de Charge dans les Sels du Tetracyanoquinodimethane (TCNQ). Acta Cryst. 1977, B33, 2744-2750. (40) Zuccaccia, C.; Bellachioma, G.; Cardaci, G.; Macchioni, A. Solution Structure Investigation of Ru(II) Complex Ion Pairs: Quantitative NOE Measurements and Determination of Average Interionic Distances. J. Am. Chem. Soc. 2001, 123, 11020-11028.

12 ACS Paragon Plus Environment

Page 13 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

(41) Macchioni, A.; Magistrato, A.; Orabona, I.; Ruffo, F.; Rothlisberger, U.; Zuccaccia, C. Direct Observation of an Equilibrium Between Two Anion-Cation Orientations in Olefin Pt(II) Complex Ion Pairs by HOESY NMR Spectroscopy. New J. Chem. 2003, 27, 455-458. (42) Neuhaus, D.; Williamson, M.: The Nuclear Overhauser Effect in Strucutral and Conformational Analysis; VCH Publishers: New York, 1989. (43) Binev, Y. I.; Georgieva, M. K.; Novkova, S. I. The Conversion of Phenylpropanedinitrile (Phenylmalononitrile) into the Carbanion, Followed by IR Spectra, Ab Initio and DFT Force Field Calculations. Spectrochim. Acta A 2003, 59, 30413052. (44) McRae, E. G. Theory of Solvent Effects on Molecular Electronic Spectra. Frequency Shifts. J. Phys. Chem. 1957, 61, 562. (45) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Structural Changes Accompanying Intramolecular Electron Transfer: Focus on Twisted Intramolecular Charge-Transfer States and Structures. Chem. Rev. 2003, 103, 3899. (46) Manohara, S. R.; Kumar, V. U.; Shivakumaraiah; Gerward, L. Estimation of Ground and Excited-State Dipole Moments of 1,2-Diazines by Solvatochromic Method and Quantum-Chemical Calculation. Mol. Liq. 2013, 181, 97. (47) Sasaki, S.; Drummen, G. P. C.; Konishi, G.-i. Recent Advanecs in Twisted Intramolecuar Charge Transfer (TICT) Fluorescence and Related Phenomena in Materials Chemisty. J. Mater. Chem. C 2016, 4, 2731-2743. (48) Kanis, D. R.; Lacroix, P. G.; Ratner, M. A.; Marks, T. J. Electronic Structure and Quadratic Hyperpolarizabilities in Organo-Metal Chromophores Having Weakly Coupled pNetworks. Unusual Mechanisms for Second-Order Response. J. Am. Chem. Soc. 1994, 116, 10089-10102. (49) Isborn, C. M.; Davidson, E. R.; Robinson, B. H. Ab Initio Diradical/Zwitterionic Polarizabilities and Hyperpolarizabilities in Twisted Double Bonds. J. Phys. Chem. A 2006, 110, 7189-7196. (50) Teran, N. B.; He, G. S.; Baev, A.; Shi, Y.; Swihart, M. T.; Prasad, P. N.; Marks, T. J.; Reynolds, J. R. Twisted ThiopheneBased Chromophores with Enhanced Intramolecular Charge Transfer for Cooperative Amplification of Third-Order Optical Nonlinearity. J. Am. Chem. Soc. 2016, 138, 6975-6984. (51) Lytel, R.; Mossman, S. M.; Kuzyk, M. G. Phase Disruption as a New Design Paradigm for Optimizing the Nonlinear-Optical Response. Opt. Lett. 2015, 40, 4735-4738. (52) Abbotto, A.; Beverina, L.; Bradamante, S.; Facchetti, A.; Klein, C.; Pagani, G. A.; Redi-Abshiro, M.; Wortmann, R. A Distinctive Example of the Cooperative Interplay of Structure and Environment in Tuning of Intramolecular Charge Transfer in Second-Order Nonlinear Optical Chromophores. Chem. Eur. J. 2003, 9, 1991-2007. (53) Chandra Ray, P. Remarkable Solvent Effects on First Hyperpolarizabilities of Zwitterionic Merocyanine Dyes: Ab Initio TD-DFT/PCM Approach. Chem. Phys. Lett. 2004, 395, 269273. (54) Hammond, S. R.; Sinness, J.; Dubbury, S.; Firestone, K. A.; Benedict, J. B.; Wawrzak, Z.; Clot, O.; Reid, P. J.; Dalton, L. R. Molecular Engineering of Nanoscale Order in Organic ElectroOptic Glasses. J. Mater. Chem. 2012, 22, 6752.

13 ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

14 ACS Paragon Plus Environment

Page 14 of 14